Collisions of the supercritical Keller-Segel particle system
Abstract.
We study a particle system naturally associated to the -dimensional Keller-Segel equation. It consists of Brownian particles in the plane, interacting through a binary attraction in , where stands for the distance between two particles. When the intensity of this attraction is greater than , this particle system explodes in finite time. We assume that and study in details what happens near explosion. There are two slightly different scenarios, depending on the values of and , here is one: at explosion, a cluster consisting of precisely particles emerges, for some deterministic depending on and . Just before explosion, there are infinitely many -ary collisions. There are also infinitely many -ary collisions before each -ary collision. And there are infinitely many binary collisions before each -ary collision. Finally, collisions of subsets of particles never occur. The other scenario is similar except that there are no -ary collisions.
Key words and phrases:
Keller-Segel equation, Stochastic particle systems, Bessel processes, Collisions2010 Mathematics Subject Classification:
60H10, 60K351. Introduction and main results
1.1. Informal definition of the model
We consider some scalar parameter and a number of particles with positions at time . Informally, we assume that the dynamics of these particles are given by the system of S.D.E.s
(1) |
where the -dimensional Brownian motions are independent. In other words, we have Brownian particles in the plane interacting through an attraction in , which is Coulombian in dimension . Actually, this S.D.E. does not clearly make sense, due to the singularity of the drift, and we will use, as suggested by Cattiaux-Pédèches [4], the theory of Dirichlet spaces, see Fukushima-Oshima-Takeda [11].
1.2. Brief motivation and informal presentation of the main results
This particle system is very natural from a physical point of view, because, as we will see, there is a tight competition between the Brownian excitation and the Coulombian attraction. It can also be seen as an approximation of the famous Keller-Segel equation [16], see also Patlak [20]. This nonlinear P.D.E. has been introduced to model the collective motion of cells, which are attracted by a chemical substance that they emit. It is well-known that a phase transition occurs: if the intensity of the attraction is small, then there exist global solutions, while if the attraction is large, the solution explodes in finite time.
We will show that this phase transition already occurs at the level of the particle system (1): there exist global (very weak) solutions if (subcritical case, see Proposition 1.3 below), but solutions must explode in finite time if (supercritical case).
To our knowledge, the supercritical case has not been studied in details, and we aim to describe precisely the explosion phenomenon. Informally, we will show the following (see Theorem 1.5 below). We assume that and , we set . There exists a (very weak) solution to (1), with a.s. and such that exists. Moreover, there is a cluster containing precisely particles in the configuration , and no cluster containing strictly more than particles. Such a cluster containing particles is inseparable, so that (1) is meaningless (even in a very weak sense) after . Just before explosion, there are infinitely many -ary collisions, where . If , we set and just before each -ary collision, there are infinitely many -collisions. Else, we set . In any case, there are infinitely many binary collisions just before each -ary collision. During the whole time interval , there are no -ary collisions, for any .
This phenomenon seems surprising and original, in particular because of the gap between binary and -ary collisions.
1.3. Sets of configurations
We introduce, for all and all ,
Here is the cardinal of and stands for the Euclidean norm in . Observe that if and only if all the particles indexed in are at the same place. We also set, for ,
which represents the set of configurations with no cluster of (or more) particles. Observe that for all .
1.4. Bessel processes
We recall that a squared Bessel process of dimension is a nonnegative solution, killed when it reaches if , of the equation
where is a -dimensional Brownian motion. We then say that is a Bessel process of dimension . This process has the following property, see Revuz-Yor [21, Chapter XI]: if , then a.s., for all , ;
if , then a.s., is reflected infinitely often at ; if , then a.s. hits and is then killed.
1.5. Some important quantities
Consider a (possibly very weak) solution to (1). As we will see, when fixing a subset and when neglecting the interactions between the particles indexed in and the other ones, one finds that the process behaves like a squared Bessel process with dimension , where
(2) |
Similar computations already appear in Haškovec-Schmeiser [12], see also [8]. A little study, see Appendix A, see also Figure 1 and Subsection 1.8 for numerical examples, shows the following facts. For , we set .
Lemma 1.1.
Fix and such that . For , we have
(3) |
We also define , and
If and , then and it holds that ; if ; if ; if .
Figure 1. Plot of as a function of with
and with (left) and (right).
, , , ,
We thus expect that there may be some non sticky -ary collisions for , some sticky -ary collisions when , but no -ary collision for .
1.6. Generator and invariant measure
As we will see in Subsection 3.13, the S.D.E. (1) cannot have a solution in the classical sense, at least when , because the drift term cannot be integrable in time. We will thus define a solution through the theory of the Dirichlet spaces.
For and for the Lebesgue measure on , we set
(4) |
where stands for the set . Informally, the generator of the solution to (1) is given by , where for ,
(5) |
see (4) for the last equality. It is well-defined for all and -symmetric. Indeed, an integration by parts shows that
(6) |
As we will see in Proposition A.1, the measure is Radon on in the subcritical case , while it is Radon on (and not on ) in the supercritical case . This will allow us to use some results found in Fukushima-Oshima-Takeda [11] and to obtain the following existence result.
Proposition 1.2.
We fix and such that and recall that . We set and , where is a cemetery point. There exists a diffusion with values in , which is -symmetric, with regular Dirichlet space on with core defined by
and such that for all , all , the law of under has a density with respect to the Lebesgue measure on . We call such a process a -process and denote by its life-time.
We refer to Subsection B.1 for a quick summary about the notions used in this proposition: diffusion (i.e. continuous Hunt process), link between its generator, semi-group and Dirichlet space, definition of the one-point compactification topology endowing , etc. Let us mention that by definition, is absorbing, i.e. for all . Also, is a priori continuous on only for the one-point compactification topology on , which precisely means that it is continuous for the usual topology of during , and it holds that for any increasing sequence of compact subsets of such that .
As we will see in Remark 11.6, for all , under , solves (1) during , where . By the Markov property, this implies solves (1) during any open time-interval on which it does not visit .
When , we have and thus . We will easily prove the following non-explosion result, which is almost contained in Cattiaux-Pédèches [4], who treat the case where .
Proposition 1.3.
Fix and . Consider the -process introduced in Proposition 1.2. For all , we have .
When , we will see that there is explosion. Note that any collision of a set of particles makes the process leave and thus explode. However, it is not clear at all at this point that explosion is due to a precise collision: the process could explode because it tends to infinity (which is not hard to exclude) or to the boundary of with possibly many oscillations.
1.7. Main result
To avoid any confusion, let us define precisely what we call a collision.
Definition 1.4.
(i) For , we say that there is a -collision in the configuration if and if for all .
(ii) For a -valued process , we say that there is a -collision at time if there is a -collision in the configuration .
The main result of this paper is the following description of the explosion phenomenon.
Theorem 1.5.
Assume that , that and recall that , and were defined in Lemma 1.1. Consider the -process introduced in Proposition 1.2. For all , we -a.s. have the following properties: (i) is finite and exists for the usual topology of ; (ii) there is with cardinal such that there is a -collision in the configuration , and for all such that , there is no -collision in the configuration ; (iii) for all and all with cardinal , there is an infinite number of -collisions during and none of these instants of -collision is isolated; (iv) if , then for all such that and , for all instant of -collision and all , there is an infinite number of -collisions during and none of these instants of -collision is isolated; (v) for all with cardinal , there is no -collision during ; (vi) for all such that and , for all instant of -collision and all , there is an infinite number of -collisions during and none of these instants of -collision is isolated.
The condition is crucial to guarantee that . On the contrary, we impose for simplicity, because Lemma 1.1 does not hold true without this assumption. The other cases may also be studied, but we believe this is not very restrictive: is thought as very large when compared to , at least as far as the approximation of the Keller-Segel equation is concerned.
1.8. Comments
Let us mention that the very precise values of and influence the value .
(a) If and , we have , and .
(b) If and , we have and .
Let us describe informally, in the chronological order, what happens e.g. in case (b) above. We start with particles at different places. During the whole story, there is no -ary collision for . Here and there, two particles meet, they collide an infinite number of times, but manage to separate. Then at some times, we have particles close to each other and there are many binary collisions. Then, if a -th particle arrives in the same zone (and this eventually occurs), there are infinitely many -ary collisions, with infinitely many binary collisions of all possible pairs before each. These particles may manage to separate forever, or for a large time, but if a -th particle arrives in the zone (and this situation eventually occurs), then there are infinitely many -ary collisions of all the possible subsets and, finally, a -ary collision producing explosion, and the story is finished. Informally, the resulting cluster is not able to separate, because the attraction dominates the Brownian excitation, since a Bessel process of dimension is absorbed when it reaches . We hope to be able, in a future work, to propose and justify a model describing what happens after explosion.
1.9. References
In many papers about the Keller-Segel equation, the parameter is used, so that the transition at corresponds to the transition at . As already mentioned, this nonlinear P.D.E. has been introduced to model the collective motion of cells, which are attracted by a chemical substance that they emit. It describes the density of particles (cells) with position at time and writes, in the so-called parabolic-elliptic case,
(7) |
Informally, this solution should be the mean-field limit of the particle system (1) as .
We refer to the recent review paper on (7) by Arumugam-Tyagi [1]. The best existence of a global solution to (7), including all the subcritical parameters , is due to Blanchet-Dolbeault-Perthame [2]. The blow-up of solutions to (7), in the supercritical case , has been studied e.g. by Fatkullin [7] and Velasquez [24, 25]. More close to our study, Suzuki [23] has shown, still in the supercritical case, the appearance of a Dirac mass with a precise (critical) weight, at explosion. This is the equivalent, in the limit , to the fact that exists and corresponds to a -collision, for some with precise cardinal . Let us finally mention Dolbeault-Schmeiser [6], who propose a post-explosion model in the supercritical case. Concerning particle systems associated with (7), let us mention Stevens [22], who studies a physically more complete particle system with two types of particles, for cells and chemo-attractant particles, with a regularized attraction kernel. Haškovec and Schmeiser [12, 13] study a particle system closer to (1), but with, again, a regularized attraction kernel. Cattiaux-Pédèches [4], as well as [8], study the system (1) without regularization in the subcritical case: existence of a global solution to (1) has been shown in [8] when , and uniqueness of this solution has been established in [4]. Also, the theory of Dirichlet spaces has been used in [4] to build a solution to (1). Finally, the limit as to a solution of (7) is proved in [8] in the very subcritical case where , up to extraction of a subsequence. This last result has been improved by Bresch-Jabin-Wang [3], who remove the necessity of extracting a subsequence and consider the (still very subcritical) case where . Olivera-Richard-Tomasevic [18] have recently established the convergence of a smoothed version of (1), for all the subcritical cases . Informally, in view of the mean distance between particles, the regularization used in [18] is not far from being physically reasonable. There is also a related paper of Jabir-Talay-Tomasevic [14] about a one-dimensional but more complicated parabolic-parabolic model.
1.10. Originality and difficulties
To our knowledge, this is the first study of the supercritical Keller-Segel particle system near explosion. We hope that this model, which makes compete diffusion and Coulomb interactions, is very natural from a physical point of view, beyond the Keller-Segel community. The phenomenon we discovered seems surprising and original, in particular because of the gap between binary and -ary collisions. We are not aware of other works, possibly dealing with other models, showing such a behavior.
In Section 3, we give the main arguments of the proofs, with quite a high level of precision, but ignoring the technical issues. While it is rather clear, intuitively, that the process explodes in finite time when and that no -collisions may occur for , the continuity at explosion is delicate, and some rather deep arguments are required to show that that each -ary collision is preceded by many binary collisions, that each -ary collision is preceded by many -ary collisions, that explosion is preceded by many -ary collisions, and that explosion is due to the emergence of a cluster with precise size (which more or less says that a possible -ary collision would necessarily be preceded by a -collision).
Actually, the rigorous proofs are made technically much more involved than those presented in Section 3, because we have to use the theory of Dirichlet spaces. Due to the singularity of the interactions and to the occurrence of many collisions near explosion, we can unfortunately not, as already mentioned, deal at the rigorous level directly with the S.D.E. (1). We thus have to use some suitable heavy versions of some usual tools such as Itô’s formula, Girsanov’s theorem, time-change, etc.
1.11. Plan of the paper
In Section 2, we introduce some notation of constant use. In Section 3, we explain the main ideas of the proofs, with quite a high level of precision, but without speaking of the heavy technical issues related to the use of the theory of Dirichlet spaces. Section 4 is devoted to the existence of a first version of the Keller-Segel process, namely without the property that has a density, and we introduce a spherical Keller-Segel process. In Section 5, we show that the Keller-Segel process enjoys a crucial and noticeable decomposition in terms of a -dimensional Brownian motion, a squared Bessel process and a spherical process. Section 6 consists in building some smooth approximations of some indicator functions that behave well under the action of the generator . In Section 7, we make use of the Girsanov theorem to prove that when two sets of particles of a -process are not too close from each other, they behave as two independent smaller -processes. In Section 8, we study explosion and continuity (in the usual sense) at the explosion time. Section 9 is devoted to establish some parts of Theorem 1.5 for some particular ranges of values of and . Using the results of Section 7, we reduce the general study to the special cases of Section 9 and we prove, in Section 10, that the conclusions of Theorem 1.5 hold true quasi-everywhere. Finally, in Section 11, we remove the restriction quasi-everywhere and conclude the proofs of Propositions 1.2 and 1.3 and of Theorem 1.5. Appendix A contains a few elementary computations: proof of Lemma 1.1, proof that is Radon on , and study of a similar measure on a sphere. We end the paper with Appendix B, that summarizes all the notions and results about Dirichlet spaces and Hunt processes we shall use.
2. Notation
We introduce the spaces
For , we have and . We consider the (unnormalized) Lebesgue measure on , as well as, recall (4),
(1) |
We define by and by
(2) |
We have and . The orthogonal projection is given by
and we introduce defined by
(3) |
For , the projections and span are given by
where .
Finally, we introduce the natural operators defined for and by
(5) |
where and stand for the usual gradient and Laplacian in . Since , with open, and since is smooth on , we can indeed define and for all . Similarly, for and , we set
(6) |
To conclude this subsection, we note that for all , for all ,
(7) |
Indeed, it suffices to observe that setting for all , we have , and and that for , we have and .
3. Main ideas of the proofs
Here we explain the main ideas of the proofs of Proposition 1.3 and Theorem 1.5. The arguments below are completely informal. In particular, we do as if our -process was a true solution to (1) until explosion and we apply Itô’s formula without care. We always assume at least that , and , which implies that .
3.1. Existence
The existence of the -process , with values in , is an easy application of Fukushima-Oshima-Takeda [11, Theorem 7.2.1]. The only difficulty is to show that the invariant measure is a Radon on , see Proposition A.1. The process may explode, i.e. get out of any compact subset of in finite time. Observe that a typical compact subset of is of the form, for ,
3.2. Center of mass and dispersion process
One can verify, using Itô’s formula, that the center of mass is a -dimensional Brownian motion with diffusion constant , that the dispersion process is a squared Bessel process with dimension , recall (2), and that these two processes are independent.
Consequently, if , the limits and a.s. exist, and this implies that : the process cannot explode to infinity, it can only explode because it tends to the boundary of . If moreover (i.e. if ), this is sufficient to show that , since then .
3.3. Behavior of distant subsets of particles
Consider a partition of . If we neglect interactions between particles of which the indexes are not in the same subset, we have, for each , setting ,
and we recognize a -process.
During time intervals where particles indexed in different subsets are far enough from each other, we can indeed bound the interaction between those particles, so that the Girsanov theorem tells us that behave similarly, in the sense of trajectories, as independent , …, -processes.
3.4. Brownian and Bessel behaviors of isolated subsets of particles
Consider . As seen just above, during time intervals where the particles indexed in are far from all the other ones, the system behaves, in the sense of trajectories, like a -process. Hence, as seen in Subsection 3.2, behaves like a -dimensional Brownian motion with diffusion constant and behaves like a squared Bessel process of dimension , which equals , recall (2).
3.5. Continuity at explosion
Here we assume that , so that and we explain why a.s., and exists, in the usual sense of .
(a) We first show that a.s. On the event where , the squared Bessel process is defined for all times. Recall that (because ) and that a squared Bessel process with negative dimension can be defined on the whole time half-line and a.s. becomes negative in finite time. Since by definition, this contradicts the fact that . Similarly, one can show that a -process has no chance to be defined after the first hitting time of by , where : this makes the choice of the space very natural. Indeed, assume that is defined during with . Consider the maximal subset of containing and such that . Then there is such that during , the particles labeled in are far from the ones labeled outside . By Subsection 3.4, behaves like a squared Bessel process with dimension issued from . But such a process is instantaneously negative, because (since ). Since , this contradicts the fact that .
(b) We next show by reverse induction that a.s. for all with , we have
(1) |
If , exists by continuity of the (true) squared Bessel process and this implies the result. We now fix and assume that (1) holds true for all such that . We consider with : by induction assumption, either there is such that and then , or for all , . In this last case, and when and (which is the negation of (1)), there are and such that (i) upcrosses infinitely often during and (ii) for all such that , the particles indexed in are far from all the other ones (because then is small and is large for all ), so that behaves like a squared Bessel process with dimension , see Subsection 3.4. Points (i) and (ii) are in contradiction, since a squared Bessel process is continuous and thus cannot upcross infinitely often during a finite time interval.
(c) We now show that exists. Using (b) and the (random) equivalence relation on defined by if and only if , one can build a (random) partition of such that for all , and . Hence, there is such that for all , the particles labeled in are far from the ones labeled in during . As seen in Subsection 3.4, we conclude that for all , behaves like a Brownian motion during , and thus exists. Since moreover , we deduce that for all , . As a conclusion exists for all .
3.6. A spherical process
We recall that , , and were introduced in Section 2 and introduce the possibly exploding (with life-time ) process with values in , informally solving (we will also use here the theory of Dirichlet spaces), for some given and some -valued Brownian motion ,
We call such a process a -process.
3.7. Decomposition of the process
We assume that and are such and, as usual, . We consider a -dimensional Brownian with diffusion constant , a squared Bessel process with dimension killed when it hits , with life-time , and a -process , these three processes being independent. We introduce the time-change
Since (because ), since and since, roughly, the paths of are -Hölder continuous, it holds that a.s. We introduce the inverse function of .
We also set and observe that , since is -valued, and that if and only if . A fastidious but straightforward computation shows that, recalling (2),
which is well-defined during , solves (1).
This decomposition of the -process, which is noticeable in that satisfies an autonomous S.D.E. and thus is Markov, is at the basis of our analysis.
In other words, is the restriction to the time interval of a -process . Moreover, we have : if is finite, then gets out of at time , so that gets out of at time , whence ; if next , then and remains in for all times, so that remains in during , whence .
We have and for all , because is -valued. By definition of , the process cannot have any -collision. But for any with cardinal at most ,
(2) |
Moreover, as seen a few lines above, is equivalent to . In other words, since for all and since , we have
(3) |
3.8. Some special cases
Using the Girsanov theorem, see Subsection 3.4, we will manage to reduce a large part of the study to the special cases that we examine in the present subsection. Here we explain the following facts, for and with : (a) if , then a.s., and for all , all with , has infinitely many -collisions during ; (b) if (whence ), then a.s., .
We keep the same notation as in the previous subsection.
(i) We first verify that in (a), . Since , it holds that . If first , then by Subsection 3.2 and we are done. If next , then and exists by Subsection 3.5. Moreover cannot belong to by definition of and thus has its particles at the same place, i.e. : we have .
(ii) In (b), by Subsection 3.5 because implies that .
(iii) We consider, in any case, the spherical process and assume that . An Itô computation shows that for , for some -dimensional Brownian motion ,
We fix to be chosen later. During time intervals where , we thus have, for some constant ,
(4) |
where we used the Cauchy-Schwarz inequality and that is uniformly bounded (because is -valued). Hence, still during time intervals where , by comparison, is smaller than , the solution to
(5) |
And a little study involving scale functions/speed measures shows that this process hits zero in finite time if and only if , exactly as a squared Bessel process with dimension .
(iv) We end the proof of (a). In this case, , so that is non-exploding, as seen in Subsection 3.6. Hence and we can use (iii). Moreover, is recurrent, still by Subsection 3.6. We fix with and we choose small enough so that we have
where is the invariant measure (1) of . Hence the process visits the zone infinitely often and each time, has a (uniformly) positive probability to hit by (iii) and since . Consequently, for any , has infinitely many -collisions during . Recalling (2) and that by (i), we conclude that for any , has infinitely many -collisions during .
(v) We finally complete the proof of (b). By (3), it is sufficient to show that a.s.
Assume that is recurrent (and thus non-exploding). Then we take and apply the same reasoning as in (iv): since , hits zero in finite time and this makes get out of and thus explode, since is -valued and since . We thus have a contradiction.
Hence is transient and it eventually gets out of the compact of
for any fixed . Hence on the event where , a.s. Recalling now that and that is -valued (whence ) we can a.s. find with such that but . It is then not too hard to find and such that each time (which often happens), all the particles indexed in are far from all the other ones with a distance greater than . We conclude from (iii), since (because ) that each time , it has a (uniformly) positive probability to hit zero. On the event , this will eventually happen, so that the process will have a -collision and thus will leave in finite time. Hence will explode, so that .
3.9. Size of the cluster
We assume that . Hence and exists, by Subsection 3.5. Moreover, by definition of , we know that . We want now to show that , i.e. that the cluster causing explosion is precisely composed of particles. If , there is nothing to do, since then . Now if , we assume by contradiction, that there is with such that and . Then there is such that during , the particles indexed in are far from the other ones, so that behaves like a -process by Subsection 3.3. Observe now that because and because . We thus know from the special case (b) of Subsection 3.8 that , which contradicts the fact that .
3.10. Collisions before explosion
We fix again . We recall that and we show that there are infinitely many -ary collisions just before explosion. We know from the previous subsection that there exists such that and and . Then there is such that during , the particles indexed in are far from the other ones, so that behaves like a -process by Subsection 3.3. Observe now that thanks to Lemma 1.1 and that because . We thus know from the special case (a) of Subsection 3.8 that has infinitely many -collisions just before , for all .
3.11. Absence of other collisions
We want to show that when , for with , there is no -collision during . Suppose by contradiction that there is with and such that and for all , . Then there is such that during , the particles indexed in are far from the other ones, so that behaves like a squared Bessel process with dimension , see Subsection 3.4. Since because , see Lemma 1.1, such a Bessel process cannot hit zero, whence a contradiction.
3.12. Binary collisions
We still assume that , we suppose that there is a -collision for some such that at some time and we want to show that there are infinitely many binary collisions just before . There is such that the particles indexed in are far from all the other ones during , so that Subsection 3.3 tells us that behaves like a -process. We observe that , that and that by Lemma 1.1.
We are reduced to show that a -process, that we still denote by , such that , and , a.s. has infinitely many binary collisions before the first instant of -collision. Such a process does not explode, because (since ), see Subsection 3.2. Hence using (2) (which is licit since ), we only have to show that e.g. collides infinitely often with during .
First, one easily gets convinced that the probability that e.g. collides with before is positive, because the probability that all the particles are pairwise far from each other, except and , during the time interval , is positive. On this kind of event, by Subsection 3.4, behaves like a squared Bessel process with dimension and thus hits zero during (and thus before ) with positive probability.
3.13. Non-integrability of the drift term
Here we check that when , the S.D.E. (1) cannot have a solution in the classical sense, because the drift term is not integrable in time. More precisely, recall that there is some -collision at some time strictly before explosion, for some with cardinal . We now show that a.s., for ,
which indeed shows the non-integrability of the drift term. Since is an instant of -collision, there exists small enough so that during , the particles labeled in are far from the particles labeled in . It clearly suffices to show that a.s., where
But
so that a.s. belongs to , and where
for each . Since the invariant measure of satisfies , it a.s. holds true that for a.e. (at least for a.e. initial condition), so that a.s., is well-defined for a.e. . We now show that is bounded from below on . We have
Using now the Cauchy-Schwarz inequality and the fact that , we find that
To conclude that a.s., it remains to verify that a.s. By Subsection 3.4, behaves like a squared Bessel process with dimension during . Since and , we conclude that indeed, a.s.: this can be shown by comparison with the -dimensional Brownian motion.
4. Construction of the Keller-Segel particle system
The aim of this section is to build a first version of the Keller-Segel particle system using the book of Fukushima-Oshima-Takeda [11]. We also build a -valued process for later use.
Proposition 4.1.
We fix and such that , recall that and that and were defined in (4) and (1). We set and , as well as and , where is a cemetery point. (i) There exists a unique diffusion with values in , which is -symmetric, with regular Dirichlet space on with core defined by
We call such a process a -process and denote by its life-time.
(ii) There exists a unique diffusion with values in , which is -symmetric, with regular Dirichlet space on with core defined by
We call such a process a -process and denote by its life-time.
The proof that we can build a -process, i.e. a -process such that has density for all and all will be handled in Section 11.
We refer to Subsection B.1 for some explanations about the notions used in this proposition: link between a diffusion (i.e. a continuous Hunt process), its generator, semi-group and its Dirichlet space, definition of the one-point compactification topology, i.e. the topology endowing and , and about the quasi-everywhere notion. The state is absorbing, i.e. for all and for all .
Remark 4.2.
By definition of the one-point compactification topology, for any increasing sequence of compact subsets of such that , . Similarly, for any increasing sequence of compact subsets of such that , .
The uniqueness stated e.g. in Proposition 4.1-(i) has to be understood in the following sense, see [11, Theorem 4.2.8 p 167]: if we have another diffusion enjoying the same properties, then quasi-everywhere, the law of under equals the law of under . The quasi-everywhere notion depends on the Hunt process under consideration but, as recalled in Subsection B.1, two Hunt processes with the same Dirichlet space share the same quasi-everywhere notion.
Proof of Proposition 4.1.
We start with (i). We consider the bilinear form on defined by . It is well-defined, since is Radon on by Proposition A.1.
We first show that it is closable, see [11, page 2], i.e. that if is such that in and , then : since is a Cauchy sequence in , it converges to a limit and it suffices to prove that a.e. For , we have . But, recalling (4),
Thus by the Cauchy-Schwarz inequality,
which tends to since in , since and since is smooth and positive on . Thus for all , so that a.e.
We can thus consider the extension of to , where we have set for .
Next, is obviously regular with core , see [11, page 6], because is dense in for the norm associated to by definition of and is dense, for the uniform norm, in . It is also strongly local, see [11, page 6], i.e. if and if is constant on a neighborhood of .
Then [11, Theorems 7.2.2 page 380 and 4.2.8 page 167] imply the existence and uniqueness of a Hunt process with values in , which is -symmetric, of which the Dirichlet space is , and such that is -a.s. continuous on for all , where .
Furthermore, since is strongly local, we know from [11, Theorem 4.5.3 page 186] that we can choose (modifying only on a properly exceptional set) such that for all . This implies that for all , -a.s., the map is continuous from to , endowed with the one-point compactification topology on recalled in Subsection B.1. Hence is a diffusion.
For (ii), the very same strategy applies. The only difference is the integration by parts to be used for the closability: for and , it classically holds that
(1) |
This can be shown naively using Lemma A.2. ∎
We now make explicit the generators of and when applied to some functions enjoying a few properties. See Subsection B.1 for a precise definition of the generator of a Hunt process. We have to introduce a few notation. For , and , we set
(2) |
where
This is in accordance with (4), in the sense that . The formula (2) makes sense for when (with replaced by ) and we recall that for and , was defined in (5) by . We will often use that for all , all , all ,
(3) |
For , and , we set
(4) |
This formula makes sense for when (with replaced by ) and we set, for and , .
Remark 4.3.
(i) Denote by the generator of the process of Proposition 4.1-(i). If satisfies , then and .
(ii) Denote by the generator of the process of Proposition 4.1-(ii). If satisfies , then and .
Proof.
To check (i), it suffices by (1) to verify that (a) , (b) and (c) for all , we have .
Point (a) is clear, since . Point (b) follows from the facts that is Radon on , that is compactly supported in and that , because for all , . Concerning (c) it suffices, by definition of and since , to show that for all , we have . But for , by a standard integration by parts, since and are smooth,
We conclude letting by dominated convergence, since and a.e., since by assumption, for some constant and for Supp which is compact in , and since .
We end the section with a quick irreducibility/recurrence/transience study of the spherical process, see Subsection B.1 again for definitions.
Lemma 4.4.
We fix and such that and consider the process and its Dirichlet space as in Proposition 4.1-(ii).
(i) is irreducible and we have the alternative: either is recurrent and in particular it is non-exploding and for all measurable such that , quasi-everywhere; or is transient and in particular for all compact set of , we have quasi-everywhere . (ii) If , then is recurrent.
In the transient case, one might also prove that , but this would be useless for our purpose.
Proof.
We start with (i). We first show that in any case, is irreducible. By [11, Corollary 4.6.4 page 195] and since with bounded from below by a constant (on ), it suffices to prove that the -symmetric Hunt process with regular Dirichlet space on with core such that for all , is irreducible. But this Hunt process is nothing but a -valued Brownian motion. This Brownian motion is a priori killed when it gets out of , but this does a.s. never occur since such a Brownian motion never has two (bi-dimensional) coordinates equal. This -valued Brownian motion is of course irreducible. We conclude from [11, Lemma 1.6.4 page 55] that is either recurrent or transient.
When is recurrent, [11, Theorem 4.7.1-(iii) page 202] gives us the result.
When is transient, we fix a compact set of and we know from Lemma A.3 that , so that by definition of transience, for -a.e , . Setting , we get in particular that for -a.e , . But, by [11, (4.1.9) page 155], is finely continuous. Using [11, Lemma 4.1.5 page 155], we deduce that quasi-everywhere. The Markov property allows us to conclude.
Concerning (ii), we recall from Proposition A.3 that , because implies that , see Lemma 1.1. Moreover, implies that , whence is compact: the process cannot explode, i.e. . Consequently, is recurrent, since belongs to and since . Indeed, as recalled Subsection B.1, if was transient, we would have for all , with the convention that . ∎
5. Decomposition
The goal of this section is to prove the following decomposition of the Keller-Segel particle system defined in Proposition 4.1-(i). This decomposition is noticeable and crucial for our purpose.
Proposition 5.1.
We fix and such that , and we recall that , that and that .
For , we set , and and we consider three independent processes: , a -dimensional Brownian motion with diffusion constant starting from , a squared Bessel process with dimension starting from and killed when it gets out of , with life-time , , a -process starting from , with life-time . We introduce , and its generalized inverse . We define , where we recall from (2) that when and where we set when or . Observe that the life-time of equals . Consider also a -process , with life-time , and , where and where . In other words, is the version of killed when it gets out of . The life-time of is .
The law of is the same as that of under , quasi-everywhere in .
We take the convention that , so that . Since and for all , Proposition 5.1 in particular implies that and are some independent squared Bessel process and Brownian motion until the first time vanishes. This actually holds true until explosion, as shown in Lemma 5.2 below. The quasi-everywhere notion refers to the Hunt process . Observe that when , we have , so that and .
Proof.
We slice the proof in several steps. The two first steps are more or less classical, even if we give all the details: we determine the Dirichlet spaces of the three processes , and involved in the construction of ; then we compute the Dirichlet space of ; we next identify the Dirichlet space of , which allows us to find the one of by a second time-change; by concatenation, we deduce the Dirichlet space of . The main computations are handled in Steps 3 and 4, where we find the Dirichlet space of , which allows us to conclude in Step 5 by uniqueness.
Step 1. First, take as in Proposition 4.1-(ii).
Second, consider a -dimensional Brownian motion with diffusion constant . We know from [11, Example 4.2.1 page 167] that is a -symmetric (here is the Lebesgue measure on ) diffusion with regular Dirichlet space on with core and for all ,
(1) |
Finally, let be a squared Bessel process of dimension killed when it gets out of and set , see Revuz-Yor [21, page 443]. Fukushima [10, Theorem 3.3] tells us that is a -symmetric diffusion (here is the Lebesgue measure on ) with regular Dirichlet space on with core where for all ,
(2) |
Together with [10, Theorem 3.3], this uses that the scale function and the speed measure of are respectively and . Actually, we don’t take the speed measure as reference measure but which is the same up to a constant.
Step 2. We apply Lemma B.3 to with , i.e. with thanks to the convention and recall that is its generalized inverse: we find that setting ,
is a -symmetric -valued diffusion with regular Dirichlet space on with core such that for all ,
(3) |
We use Lemma B.5 and the notation therein: recalling that , with the convention that , and that if and , it holds that
is a -symmetric -valued diffusion with regular Dirichlet space given by on with core , and for all ,
We now apply Lemma B.3 to with for all and all . We consider the time-change , with the convention that as soon as . We also set . As we will see in a few lines, it holds that
(4) |
Hence Lemma B.3 tells us that
is a -symmetric -valued diffusion with Dirichlet space on , regular with core and for all ,
(5) |
We now check the claim (4). Recall that explodes at time , that and that is the generalized inverse of . Hence is the true inverse of and we have , whence for . We also have for . Next, for , because if , i.e. if . Hence , the generalized inverse of , equals during , thus in particular for . As conclusion, (4) holds true for . If now , then , because is the generalized inverse of and because for all ,
Hence, still if , we have , while because either and thus or and thus so that . We have proved (4).
We finally conclude, thanks to Lemma B.5 again, setting with the convention that and setting in the case where and , that
is a -symmetric -valued diffusion with regular Dirichlet space on , with core . Moreover, for all ,
(6) |
For the second line, we used (5). For the last line, we used (1), (3) and the expression of , see Proposition 4.1-(ii).
Step 3. We recall that , where for and for . One easily checks that is a bijection from to , recall that and .
We now study
where , and for .
First, is a -valued diffusion, because the bijection from to is continuous, both sets being endowed with the one-point compactification topology, see Subsection B.1.
Next, we prove that is -symmetric: if are nonnegative measurable functions on and , we have, thanks to Lemma A.2 (recall that ),
But , so that
Using that is -symmetric and then the same computation in reverse order, one concludes that as desired.
Thus has a Dirichlet space on that we now determine. For , using as above Lemma A.2 and that ,
Since is bijective, we deduce, see [11, Lemma 1.3.4 page 23], that
(7) | |||
(8) |
We recall that for , we call the total gradient of at , and we have for each . And for , where is open in , we denote by the differential of at .
We start with the study of , where we recall that was introduced in Section 2 and that is defined on a neighborhood of in , see (3). It holds that for all and all , and ,
For the first equality, it suffices to use that is linear, so that . The second equality is obvious. For the third equality, which is the differential at of the function defined for (which is open in and contains ), we write . But , where , and we have and for . All in all, .
First, we have . Indeed, for all , it holds that
which, by definition of , equals .
This implies that
(10) |
Indeed, recalling the expression of , see Section 2, it suffices to note that for all , .
Next, . Indeed, for ,
which is nothing but .
This implies, recalling that is the orthogonal projection on , that
(11) |
since , so that and .
This implies that
(12) |
Step 5. As a last technical step, we verify that is a regular Dirichlet space on with core , i.e. that for all , there is such that .
Recalling (7) and using that on is regular with core , there is such that
Setting , it holds that and we have, by (8),
as well as, by Lemma A.2,
Step 6. By Steps 3, 4 and 5, we know that is a -symmetric -valued diffusion with regular Dirichlet space with core and with for .
Actually, and are some independent squared Bessel process and Brownian motion until explosion (and not only until the first time where , as shown in Proposition 5.1), a fact that we shall often use.
Lemma 5.2.
We fix and such that and we consider a -process . Quasi-everywhere, there are a -Brownian motion with diffusion constant issued from and a squared Bessel process with dimension issued from (killed when it gets out of if ) independent of such that -a.s., and for all .
Proof.
If , this follows from Proposition 5.1: setting , we have . Indeed, on , we have , whence since with (because ), which contradicts the fact that .
We now suppose that , so that and thus . We introduce the shortened notation , and split the proof in three parts.
Step 1. First, one can show similarly (but much more easily) as in the proof of Proposition 5.1 that there exists a -Brownian motion independent of , such that for all . This moreover shows that is independent of , because .
Step 2. We consider some function such that on and . Such a function exists by Remark 6.3. For , we set and show that and that for all ,
(13) |
To this end, we apply Remark 4.3. Since and since , we have to show that , and we will deduce that . By (3), we have . The only difficulty consists in showing that . Using that , we find Hence by symmetry,
(14) |
Besides, whence
(15) |
We conclude by combining (5) and (15) that
We immediately deduce, since is compactly supported, that , whence . Hence and . Moreover, recalling that with and that on , we conclude that for , whence (13), because .
Step 3. We define . By Lemma B.2 and Step 1, for all , quasi-everywhere in , is a -martingale. Recalling (13), we classically conclude that there is a Brownian motion such that during . We recognize the S.D.E. of a squared Bessel process with dimension , see Revuz-Yor [21, Chapter XI]. Since we know from Remark 4.2 that , the proof is complete. ∎
6. Some cutoff functions
We will need several times to approximate some indicator functions by some smooth functions, on which the generator (or ) is bounded. This does not seem obvious, due to the singularity of . We recall that and were defined in (2) and (4).
Lemma 6.1.
Fix , , recall that and that . Consider a partition and define, for , (with the convention that ),
(i) For all , there is a family of open relatively compact subsets of such that
and some of -valued functions such that for some , for all ,
(ii) With the same sets as in (i), there is a family of functions with values in such that for all ,
The section is devoted to the proof of this lemma. We start with the following technical result.
Lemma 6.2.
We define the family by and for all , . For all , all , all such that
it holds that for all , all .
Proof.
We fix , and as in the statement and assume by contradiction that there are , such that . Then for all ,
This implies that
whence
which is a contradiction. ∎
We are now ready to give the
Proof of Lemma 6.1.
We introduce some nondecreasing function such that on and on . We divide the proof in three steps.
Step 1. We fix and define, for , using the family of Lemma 6.2,
where . We have
(1) |
Since implies that for all , we also have
(2) |
We now show, and this is the main difficulty of the step, that for all , all with , we have . Since , we only have to verify that , where
with
Using that , we now write
where,
We have because is bounded and because
Next, we assume that (else ) and observe that implies that (because on ) and that (because on ). By Lemma 6.2, this implies that . We immediately conclude that .
Step 2. We can now prove (i). We fix and a partition of . For some to be chosen later (as a function of ), for each , we set
where with the extension .
First, is clearly included in and relatively compact in . We deduce from (2) that, setting ,
By (2) again, we can choose large enough so that contains . Next, by (1), it holds that , that on and that
which is compact in . Moreover, . Since there exists such that , we conclude that Supp .
It remains to show that . Introducing
which belongs to by Step 1, we have (with the chosen value of ) and thus by (3)
The first term is uniformly bounded because is bounded and supported in and because by Step 1 and (3). The third term is also uniformly bounded, since and since is bounded and supported in . Finally, the middle term is bounded because is bounded by and because is uniformly bounded, as we now show: is obviously bounded since and, since ,
This last quantity is uniformly bounded, since is bounded and vanishes on .
Step 3. We now prove (ii), by showing that the restriction satisfies the required conditions. We obviously have and on . It remains to show that , recall (4). Since , is bounded. We thus only have to verify that , where
Setting and using (7),
Since now and since and are self-adjoint, as every orthogonal projection, we get
But is uniformly bounded by point (i) and since is bounded on . Next, is smooth and thus bounded on . Finally,
is also uniformly bounded. ∎
Remark 6.3.
We have proved in Step 2 that for each , satisfies on and .
7. A Girsanov theorem for the Keller-Segel particle system.
In this section, we prove a rigorous version of the intuitive argument presented in Subsection 3.4.
For , all , we denote by . For a partition of , for , we abusively denote by the element of such that for all , .
We adopt the convention that for any , a -process is a -dimensional Brownian motion. This is natural in view of (1).
Proposition 7.1.
Let , such that and set . Fix some partition of with . Consider the state spaces and, for each ,
Consider
a -process,
For all , a -process.
We set and , with the convention that as soon as for some . We also introduce , as well as for all .
We fix , recall that
and set
Fix . Quasi-everywhere in , there is a probability measure on , equivalent to , such that the law of the process under is the same as that of on under .
Furthermore, the Radon-Nikodym density is -measurable, where as usual , and there is a deterministic constant such that quasi-everywhere in ,
The quasi-everywhere notion refers to the process . Let us mention that for the life-time of , we have when because . Although this is not clear at this point of the paper, the event has a positive probability if .
Proof.
We only consider the case where . The general case is heavier in terms of notation but contains no additional difficulty. We fix a non-trivial partition of . The main idea is to apply Lemma B.7 to with the function
(1) |
Unfortunately, this is not licit because .
Step 1. Set and fix and . We first compute the Dirichlet space of killed when it gets outside of , recall Lemma 6.1. Consider the measures
on and , with if . Recall that , see (4) and that by definition, see (1), : we deduce that
By Proposition 4.1, for , is a -valued -symmetric (since ) diffusion with regular Dirichlet space with core and, for , . This also holds true if e.g. , see [11, Example 4.2.1 page 167], since then is nothing but the Lebesgue measure on . Since now , by Lemma B.5, is a -symmetric -valued diffusion with regular Dirichlet space on with core and, for ,
Finally, we apply Lemma B.6 to with the open set , to find that the resulting killed process
is a -symmetric -valued diffusion with regular Dirichlet space with core such that for all ,
Step 2. We now fix and introduce, for each , , recall (1) and Lemma 6.1, and . We check here that the functions and satisfy the assumptions of Lemma B.7 (to be applied to ), that and that
(2) |
First, because , and is bounded, uniformly in , because is bounded by and vanishes outside (see Lemma 6.1), while is smooth on . To show that , it suffices by Remark 4.3 to verify that , which is clear, and that . We have
Since , the only difficulty is to check that . By (3),
Again, the only difficulty consists of the first term, because is uniformly bounded by Lemma 6.1 and vanishes outside , while is smooth on . Since Supp , we are reduced to show that . But
and we only have to verify that .
For and , we have
Hence , where
and (resp. ) is defined as (resp. ) exchanging the roles of and . First, (and ) is obviously uniformly bounded on . Next, by symmetry,
Moreover, there is such that for all , all such that , all ,
so that (and ) is bounded on , uniformly in , as desired. Finally, the above computations, together with the facts that on , also show that for ,
which is bounded on . Since , this implies that and completes the step.
Step 3. We apply Lemma B.7 to the process with and defined in Step 2. Recalling that and using the conventions and , we set
(3) |
Set . By Lemma B.7, there is a family of probability measures such that
for all and quasi-everywhere in , and such that
is a -symmetric -valued diffusion with regular Dirichlet space with core such that for all ,
Next, we apply Lemma B.6 to with the open set : the resulting killed process
is a -symmetric -valued diffusion with regular Dirichlet space with core such that for all ,
Comparing this Dirichlet space with the one found in Step 1, using that on and a uniqueness argument, see [11, Theorem 4.2.8 p 167], we conclude that quasi-everywhere in , the law of under equals the law of under .
Step 4. We fix and and complete the proof. Since on , we know from Step 3 that for all , quasi-everywhere in , for all continuous bounded , (observe that )
where and . Since is a -martingale by Lemma B.7, we deduce that quasi-everywhere in ,
(4) |
Recall that , see Lemma 6.1. Hence , , and for each , there is such that for all . We deduce from (4) that quasi-everywhere in , the process is a -martingale under . Moreover, recalling the expression (3) of , that and the bound (2), we conclude that there is a constant such that quasi-everywhere in ,
Hence the martingale is closed by some -measurable random variable that satisfies , and (4) implies that for all ,
Letting , we find that quasi-everywhere in , for ,
Setting completes the proof. ∎
8. Explosion and continuity at explosion
In this section we consider a -process with life-time . We show that when and that when . In the latter case, we also prove that a.s. exists, for the usual topology of : the Keller-Segel process is continuous at explosion. This is not clear at all at first sight: we know that a.s. for the one-point compactification topology, which means that the process escapes from every compact of , but it could either go to infinity, which is not difficult to exclude, or it could tend to the boundary of without converging, e.g. because it could alternate very fast between having its particles labeled in very close and having its particles labeled in very close. The goal of the section is to prove the following result.
Proposition 8.1.
Fix and such that , set and and consider a -process with life-time . (i) If , then quasi-everywhere, . (ii) If , then quasi-everywhere, -a.s., and exists for the usual topology of and does not belong to .
We first show that the process does not explode in the subcritical case and cannot go to infinity at explosion in the supercritical case.
Lemma 8.2.
(i) If and , then quasi-everywhere, . (ii) If and , then quasi-everywhere,
Proof.
The arguments below only apply quasi-everywhere, since we use Proposition 5.1. In both cases, we have for all and all ,
By Lemma 5.2, there are a Brownian motion and a squared Bessel process with dimension (killed when it gets out of if ), such that and for all . These processes being locally bounded, we conclude that
(1) |
(i) When and , we have , so that . Hence on the event , we necessarily have , and this is incompatible with (1) with .
(ii) When , we have , so that is killed at some finite time . It holds that . Indeed, on the event where , we have , so that (since ), which is not possible since . Hence is also a.s. finite and it holds that a.s. by (1) with the choice . ∎
To show the continuity at explosion in the supercritical case, we need to prove the following delicate lemma.
Lemma 8.3.
Assume that . Quasi-everywhere, for all with ,
Proof.
We proceed by reverse induction on the cardinal of . If first , the result is clear because is a (killed) squared Bessel process on by Lemma 5.2 (and since exactly as in the proof of Lemma 8.2-(ii)), hence it has a limit in as . Then, we assume that the property is proved if where , we take such that and we show in several steps that a.s., either or .
Step 1. We fix and introduce and, for ,
with the convention that . We show in this step that for all deterministic , there exists a constant such that for all , quasi-everywhere, on ,
where , and where, setting (recall Lemma 6.2),
By the strong Markov property of , on ,
where
and
We used that on by definition of , so that under . Using again that on , it suffices to show that there is a constant such that quasi-everywhere in .
If first or , then clearly, .
Otherwise, , where
as in Proposition 7.1 with , because and because and imply that for all , by Lemma 6.2. For the very same reasons and by definition of , it holds that
(2) |
We now apply Proposition 7.1 with (and ) and we find that quasi-everywhere in ,
(3) |
But we know from Proposition 7.1 and Lemma 5.2 that under , is a squared Bessel process with dimension , issued from , stopped at time , where . Hence there exists, under , a squared Bessel process with dimension such that for all . We introduce and we observe that
Indeed, we used that on , we have by (2) so that for all , from which we conclude that if and only . Coming back to (3), we get
The step is complete, since is the probability that a squared Bessel process with dimension issued from remains below during and is thus strictly positive, uniformly in (such that and ).
Step 2. We prove here that for all , all , quasi-everywhere,
All the arguments below only hold quasi-everywhere, even if we do not mention it explicitly during this step. For , we introduce, with defined in Step 1,
and we first show that . To this end, it suffices to check that for all , . Since is -measurable, for all ,
Since moreover and since , we deduce that on ,
so that by Step 1. Hence we conclude that
for all , so that as desired.
Hence , so that a.s., an infinite number of are realized. Recalling that
we find the following alternative:
either there is such that ; or for all , and for infinitely many ’s, which implies that because necessarily, by definition of the sequence and by continuity of on ; or for all , and there are infinitely many ’s for which and this implies that , because as previously.
Step 3. We conclude. Applying Step 2, we find that quasi-everywhere, -a.s., for all and all ,
By Lemma 8.2-(ii), we know that , so that choosing , we conclude that quasi-everywhere, -a.s., for all
(4) |
And by Lemma 8.2-(ii) again, for some (random) .
On the event where , there exists some (random) such that , whence by induction assumption, and this obviously implies that .
On the complementary event, we fix such that and we conclude from (4) and the fact that that for all , there exists such that . Recalling the definition of , we deduce that for all , upcrosses the segment a finite number of times during . Hence for all , there exists such that either for all or for all . If there is such that for all , then . If next for all , we have for all , then .
Hence in any case, we have either or . ∎
We finally give the
Proof of Proposition 8.1.
Point (i), which concerns the subcritical case, has already been checked in Lemma 8.2-(i). Concerning point (ii), which concerns the supercritical case , we already know that quasi-everywhere, by Lemma 8.2-(ii), and it remains to prove that -a.s., exists and does not belong to . We divide the proof in four steps.
Step 1. For a partition of and , we consider as in Proposition 7.1
and . We show here for each , quasi-everywhere in , -a.s., for all , all , has a limit in as .
If , the result is obvious since is a Brownian motion during by Lemma 5.2. If next , Proposition 7.1 and Lemma 5.2 tell us that under , which is equivalent to , the processes are some Brownian motions on , and thus have some limits as .
Step 2. For and a partition of , we set and, for ,
with the convention that . Using Step 1 and the strong Markov property, we conclude that quasi-everywhere, -a.s., for all , all , all , on , for all , admits a limit in as goes to . Choosing , we conclude that quasi-everywhere, -a.s., on , for all , all , all ,
Step 3. We now check that quasi-everywhere, -a.s., there is a partition of , some and some such that (i) and and (ii) for all , .
By Lemma 8.3, we know that for all , we have the alternative or . Hence the partition of consisting of the classes of the equivalence relation defined by if and only if satisfies that for all , and .
Using moreover that according to Lemma 8.2, we deduce that there is and such that for all , belongs to . Finally, we consider , which is finite by continuity of on , and it holds that and that .
Step 4. We consider the (random) partition introduced in Step 3. By Step 2 and since and , we know that quasi-everywhere, -a.s., for all , exists in . By Step 3, we know that for all , . We easily conclude that quasi-everywhere, -a.s., for all , all , . This shows that quasi-everywhere, -a.s., exists in . Moreover, cannot belong to , because when is endowed with the one-point compactification topology, see Subsection B.1. ∎
9. Some special cases
During a -collision, the particles labeled in are isolated from the other ones. Thanks to Proposition 7.1, it will thus be possible to describe what happens in a neighborhood of the instant of this -collision, by studying a -process. In other words, we may assume that , so that the following special cases, which are the purpose of this section, will be crucial.
Proposition 9.1.
Let and such that . Consider a -process as in Proposition 4.1. Recall that and set with the convention that , so that .
(i) If and , then quasi-everywhere,
(ii) If and , then quasi-everywhere, -a.s, for all with cardinal , there is such that .
(iii) If , then quasi-everywhere, -a.s, for all with cardinal , there is such that .
The proof of this proposition is very long. First, we recall some notation about the decomposition of obtained in Proposition 5.1 and we study the involved time-change. We then derive a formula describing , valid on certain time intervals, for any . This formula is of course not closed, but it allows us to compare , when it is close to , to some process resembling a squared Bessel process, of which one easily studies the behavior near . Finally, we prove Proposition 9.1, unifying a little points (i) and (ii) and treating separately point (iii).
9.1. Notation and preliminaries
We recall the decomposition of Proposition 5.1, which holds true quasi-everywhere in . Consider a Brownian motion with diffusion coefficient starting from , a squared Bessel process starting from killed when leaving with life-time and a -process starting from with life-time , all these processes being independent. For , we put . We also consider the inverse of .
Lemma 9.2.
If , then and a.s.
Proof.
Since is a (killed) squared Bessel process with dimension , we have a.s according to Revuz-Yor [21, Chapter XI]. Moreover, there is a Brownian motion such that for all , where . A simple computation shows the existence of a Brownian motion such that for all ,
Hence for all , . On the event where , we have . Hence a.s. ∎
From now on, we assume that . Hence is an increasing bijection, as well as . By Proposition 5.1, quasi-everywhere in , we can find a triple as above such that for our process starting from , for all , and actually for all because since is -valued,
We recall that if and if . Observe that , where .
We note that if , then , because is an increasing bijection from into . Hence, still if , then explodes at time strictly before , whence
(1) |
Finally note that since is -valued, it cannot have a -collision. But for any with cardinal , it holds that
(2) |
which follows from the facts that
for all , if and only if ;
is an increasing bijection from into , because .
We conclude this subsection with a remark about the quasi-everywhere notions of and , in the case where they are related as above. See Subsection B.1 for a short reminder on this notion.
Remark 9.3.
Fix such that quasi-everywhere (here quasi-everywhere refers to the Hunt process ). Then quasi-everywhere (here quasi-everywhere refers to the Hunt process , which is killed when it gets outside ).
Proof.
By definition, there exists a properly exceptional set relative to such that for all , . Thus for all , .
By Proposition 5.1, there exists a properly exceptional set relative to , such that for all , the law of under is equal to the the law of under , with some obvious notation.
Hence we only have to prove that is properly exceptional for .
First, we have for all for all . Indeed, since , the law of under equals the law of under . Since for all for all and since , we have for all for all . Hence for all for all . Finally, for all for all because is properly exceptional for .
9.2. An expression of dispersion processes on the sphere
We now study the dispersion process , for . The equation below can be informally established if assuming that (1) rigorously holds true, after a time-change and several Itô computations.
Lemma 9.4.
Fix and such that and recall that . Consider a -process with life-time , fix such that , and set . Recall that was introduced in Lemma 6.1, and observe that
Quasi-everywhere in , enlarging the filtered probability space if necessary, there exists a -dimensional -Brownian motion under such that
(3) | ||||
for all , where .
As usual, because . Note also that if , then for all , and that the constant process indeed solves (3).
Proof.
We divide the proof in several steps. The main idea is to compute and and to use that , for some martingale of which we can compute the bracket. However, we need to regularize and to localize space in a zone where the last term of (3) is bounded.
Step 1. We fix and and recall , compactly supported in , was defined in Lemma 6.1. We want to apply Remark 4.3 to and . We thus have to show that and belong to for all , which is clear, and that
for all . Since
(4) |
for all and recalling that by Lemma 6.1 and that is compactly supported in , the only issue is to verify that, for compact in ,
(5) |
Step 2. Here we prove that
(6) | ||||
and this will imply (5): the first four terms are obviously uniformly bounded on , and the last one is uniformly bounded on (because is compact in ). This will also imply, taking and observing that and , that for all ,
(7) |
Step 2.1. We first verify that for all ,
(8) | |||
(9) |
First, a simple computation shows that for , for ,
(10) |
so that in particular and
(11) |
Next, proceeding as in (7), we get for all , so that
We used that thanks to (10), that by (10) and that by (11). We first conclude that for , since and ,
(12) |
which implies (8) by (10). Second, we deduce that for ,
Using that , we conclude that for , since , and by (11),
Step 2.2. We fix and show that setting , it holds that
(13) | ||||
By (8), we may write , where
First, by symmetry,
Second, by symmetry,
Step 3. By Steps 1 and 2, we can apply Remark 4.3 and Lemma B.2: quasi-everywhere, for all , there exist two -martingales and under , such that
for all . We recall that and introduce
Since and since increases to as , see Lemma 6.1, we conclude that . Next, since on , we have, for all ,
(14) | |||
(15) |
Applying the Itô formula to compute from (14), recalling from (4) that and comparing to (15), we obtain that for ,
Hence, enlarging the probability space if necessary, we can find a Brownian motion , which is defined by for and which is then extended to , such that during . Hence, still for ,
(16) |
But by (12), whence
Inserting this, as well as the expression (7) of , in (16), shows that satisfies the desired equation on . Since a.s., the proof is complete. ∎
9.3. A squared Bessel-like process
The equation obtained in the previous lemma will be studied by comparison with the process we now introduce. This process behaves, near , like a squared Bessel processes.
Lemma 9.5.
Fix , and such that . For a -dimensional Brownian motion and for , consider the unique solution of
(17) |
For , set . For all , it holds that .
Proof.
This equation is classically well-posed, since the diffusion coefficient is -Hölder continuous and the drift coefficient is Lipschitz continuous, see Revuz-Yor [21, Theorem 3.5 page 390]. As in Karatzas-Shreve [15, (5.42) page 339], we introduce the scale function
This function is obviously continuous on and one gets convinced, for example approximating by , that it is also continuous at because . By [15, (5.61) page 344], we have
(18) |
for all . This last quantity is nonzero (which would not be the case if , since then ). ∎
9.4. Collisions of large clusters
We are now ready to give the
Proof of Proposition 9.1-(i)-(ii).
We fix , such that . We always assume that and we use the notation of Subsection 9.1. Step 1. We consider and such that and . We introduce the constant with defined in Lemma 6.2. We prove in this step that there are some constants and such that, setting
with the convention that , it holds that quasi-everywhere on ,
We introduce . We note that for all , and so that thanks to the definition of and to Lemma 6.2. This implies that , where we recall that was defined in Lemma 9.4, and that .
By the Cauchy-Schwarz inequality, and since is bounded on , there is a deterministic constant , allowed to change from line to line, such that for all , we have
where is chosen small enough so that . Actually, is only introduced to make the drift coefficient of (17) Lipschitz continuous.
Recalling that , the formula describing for , see Lemma 9.4, considering the process solution to (17) with , , and with introduced a few lines above, driven by the same Brownian motion , and using the comparison theorem, we conclude that for all .
Setting for and recalling the definition of , we conclude that Indeed, on , either , or reaches at time and we then have . In both cases, . Hence, using again that for all ,
But gives . Hence
This last quantity equals and does not depend on such that . But by Lemma 9.5 and since . Hence there exists so that and this completes the step.
Step 2. We prove (ii), i.e. that when , for any with cardinal , quasi-everywhere, -a.s., vanishes during . By (2) and Remark 9.3, and since quasi-everywhere by Lemma 4.4-(ii), it suffices to check that quasi-everywhere, -a.s., vanishes at least once during .
We fix with , set and introduce and for all ,
with defined in Step 1. All these stopping times are finite since is recurrent by Lemma 4.4-(ii). We also put, for ,
We now prove that quasi-everywhere, and this will complete the proof of (ii). For , since is -measurable, the strong Markov property tells us that
We now prove that quasi-everywhere on . For such a , we have . Moreover, for all , we have thanks to our choice of . Hence , recall Step 1. Since finally and since for all and all ,
Hence Step 1 tells us that quasi-everywhere on .
Since , we have proved that for all ,
This allows us to conclude that indeed, .
Step 3. We prove (i), i.e. that if , then quasi-everywhere. By Remark 9.3 and (1), it suffices to show that quasi-everywhere, .
For all , all , we introduce and for all ,
with defined in Step 1 and with the convention that . Step 3.1. We fix and assume that , so that by Lemma 1.1. We prove here that quasi-everywhere, -a.s., either there is such that or there is such that or there is such that . It suffices to prove that , where
But for all , is -measurable, whence, by the strong Markov property,
We used Step 1, that on the event , as well as the inclusion . One easily concludes.
Step 3.2. For all such that , quasi-everywhere, -a.s., there is no such that . Indeed, on the contrary event, there is such that , whence , which contradicts the fact that .
Step 3.3. We show by decreasing induction that
holds true for every .
The result is clear when , because for all , .
We next assume for some and we show that is true. We fix with cardinal and we apply Step 3.1 with and with some ( is random but we may apply Step 3.1 simultaneously for all ) and Step 3.2, we find that on the event , there either exists such that or such that . This second choice is not possible, since by induction assumption, for all and all . Hence there is such that .
By definition of , this implies that, still on the event where , there exists such that for all , either or . Using again the induction assumption, we get that the second choice is never possible, so that actually, for all . Since is continuous and positive on according to Step 3.2, this completes the step.
Step 3.4. We conclude from Step 3.3 that quasi-everywhere, -a.s. on the event , for all , where
This (random) set is compact in , so that Lemma 4.4-(i) tells us, both in the case where is recurrent and in the case where is transient, that this happens with probability . Hence quasi-everywhere, as desired. ∎
9.5. Binary collisions
We finally give the
Proof of Proposition 9.1-(iii).
We assume that , that and observe that and , so that and . The -process is non-exploding by Proposition 8.1-(i), and the -process is irreducible recurrent by Lemma 4.4-(ii). In particular, a.s. We divide the proof in 4 steps. First, we prove that may have some binary collisions with positive probability. Then we check that this implies that also may have some binary collisions with positive probability. Since is recurrent, it will then necessarily be a.s. subjected to (infinitely many) binary collisions. Finally, we conclude using (2).
Step 1. We set and
with large enough so that . We show in this step that quasi-everywhere in , where
To this end, we fix and introduce the set
and for . Clearly, there is some such that
where as usual , recall that because .
Since is obviously included in , we conclude that
by Proposition 7.1 with . We now set . Proposition 7.1 tells us that, quasi-everywhere in , the law of under equals the law of where is a -process issued from , where for all , is a -process, i.e. a -dimensional Brownian motion, issued from , and where all these processes are independent. We have set . This implies, together with the fact that , that
quasi-everywhere in , where
and where for all . Of course, for all , since is a Brownian motion issued from . Moreover, we know from Lemma 5.2 that is a -dimensional Brownian motion with diffusion coefficient issued from , that is a squared Bessel process of dimension issued from , and that these processes are independent. Hence, recalling the definition of ,
This last quantity is clearly positive, because a squared Bessel process with dimension , see Lemma 1.1, does hit zero, see Revuz-Yor [21, Chapter XI].
Step 2. We now deduce from Step 1 that the set is not exceptional for . Indeed, if it was exceptional, we would have quasi-everywhere. By (2) and Remark 9.3, this would imply that quasi-everywhere, , where and . But on the event defined in Step 1, there is such that and it holds that . As a conclusion, quasi-everywhere in , whence a contradiction, since .
Step 3. Since is irreducible-recurrent and since is not exceptional, we know from Fukushima-Oshima-Takeda [11, Theorem 4.7.1-(iii) page 202] that quasi-everywhere,
Step 4. Using again (2) and Remark 9.3 and recalling that and that is an increasing bijection from to , we conclude that quasi-everywhere, -a.s., visits (an infinite number of times) during . Of course, the same arguments apply when replacing by any subset of with cardinal , and the proof is complete. ∎
10. Quasi-everywhere conclusion
Here we prove that the conclusions of Theorem 1.5 hold quasi-everywhere.
Partial proof of Theorem 1.5.
We assume that and , so that , and consider a -valued -process with life-time as in Proposition 4.1, where . Preliminaries. For and , we write and instead of and with as in Proposition 7.1. We also write instead of and recall that it is equivalent to on .
Setting and , we know that quasi-everywhere in , the law of under is the same as the law of , where is a -process issued from and is a -process issued from , these two processes being independent, and where . We denote by and the life-times of and . The life-time of is given by and it holds that .
No isolated points. Here we prove that for all with , quasi-everywhere, we have , where has an isolated point and
On , we can find such that and such that there is a unique with and . By continuity, we deduce that on , there exist and such that , for all and such that has an isolated point. It thus suffices that for all and all , that we all fix from now on, quasi-everywhere, , where
By the Markov property, it suffices that quasi-everywhere in and, by equivalence, that quasi-everywhere in . We write, recalling the preliminaries,
But is a -process, so that we know from Lemma 5.2 that is a squared Bessel process with dimension . Such a process has no isolated zero, see Revuz-Yor [21, Chapter XI].
Point (i). We have already seen in Proposition 8.1-(ii) that quasi-everywhere, -a.s., and exists in and does not belong to .
Point (ii). We want to show that quasi-everywhere, -a.s., there is with such that there is a -collision and no -collision with in the configuration . We already know that , so that there is with such that there is a -collision in the configuration . Hence the goal is to verify that quasi-everywhere, for all with , , where
On , there is such that . By continuity, there also exists, still on , some such that for all . Hence we only have to prove that for all , all , all such that , quasi-everywhere, , where
By the Markov property, it suffices that quasi-everywhere in , for all and all . We now fix and . By equivalence, it suffices to prove that . Using the notation introduced in the preliminaries, we write
But is a -process with and with by Lemma 1.1 because . We also have . Hence Proposition 9.1-(i) tells us that .
Point (iii). We recall that and we fix with and . We want to prove that quasi-everywhere, -a.s., if , then for all , the set is infinite and has no isolated point. But since , see Lemma 1.1, we already know that has no isolated point. It thus suffices to check that quasi-everywhere, for all , we have , where
We used that since , for all , there is a collision in the configuration if and only if . On , thanks to point (ii) , there are , and such that and for all . Thus it suffices to prove that for all and all , that we now fix, quasi-everywhere, , where
By the Markov property, it suffices that quasi-everywhere in and, by equivalence, that . Recalling the preliminaries, we write
Setting , we observe that . Indeed, and is a -process, of which the state space is given by , where for all such that , because . Hence , so that
This last quantity equals zero by Proposition 9.1-(ii), since by Lemma 1.1 and since and since .
Point (iv). We assume that , i.e. that . We fix with and . We want to prove that quasi-everywhere, -a.s., for all , if there is a -collision in the configuration , then for all , the set is infinite and has no isolated point. We already know that has no isolated point. It thus suffices to check that quasi-everywhere, for all , we have , where
We set there is a -collision in the configuration . It holds that
On , there exists such that , so that by continuity, there exists such that for all . Observe that and that for all , there is a -collision at time if and only if , by definition of and since . All in all, it suffices to prove that for all , all , all , quasi-everywhere, where
By the Markov property, it suffices to prove that quasi-everywhere in and, by equivalence, we may use instead of . But recalling the preliminaries,
where we have set . Finally, by Proposition 9.1-(ii), because is a -process, because , because and because .
Point (v). We fix with cardinal , so that . We want to prove that quasi-everywhere, -a.s., for all , there is no -collision in the configuration . We introduce there is a -collision in the configuration , with the convention that , and we have to verify that quasi-everywhere, .
On the event , there exist and such that for all . Hence it suffices to check that for all , all and all , which we now fix, quasi-everywhere, , where
By the Markov property, it suffices that quasi-everywhere in and, by equivalence, that . Recalling the preliminaries, we write
where we have set there is a -collision in the configuration . Since is a -process, we know from Lemma 5.2 that is a squared Bessel process with dimension . Such a process does a.s. never reach .
11. Extension to all initial conditions in
We first prove Proposition 1.2: we can build a -process, i.e. a -process such that is absolutely continuous for all and all . We next conclude the proofs of Proposition 1.3 and of Theorem 1.5.
11.1. Construction of a -process
We fix and such that during the whole subsection. For each , we introduce such that for all and we set, for ,
We then consider the -valued S.D.E
(1) |
which is strongly well-posed, for every initial condition, since the drift coefficient is smooth and bounded. We denote by the corresponding Markov process.
Lemma 11.1.
For all , is a -symmetric -valued diffusion with regular Dirichlet space with core such that for all ,
Moreover has a density with respect to the Lebesgue measure on for all and all .
Proof.
Classically, is a -symmetric diffusion and its (strong) generator satisfies that for all , all , . Hence, see Subsection B.1, one easily shows that for the Dirichlet space of , we have and, for , . Since is closed, we deduce that
where . But thanks to [11, Lemma 3.3.5 page 136],
where is understood in the sense of distributions. Since finally
has the announced Dirichlet space. Finally, the absolute continuity of , for and , immediately follows from the (standard) Girsanov theorem, since the drift coefficient is bounded. ∎
For all we set . For , we introduce the open set
(2) |
We also fix a -process for the whole subsection.
Lemma 11.2.
There exists an exceptional set with respect to such that for all , for all , the law of under equals the law of under , where
Proof.
We fix . Applying Lemma B.6 to and with the open set , using that on and Lemma 11.1, we find that the processes and killed when leaving have the same Dirichlet space. By uniqueness, see [11, Theorem 4.2.8 page 167], there exists an exceptional set such that for all , the law of killed when leaving under equals the law of killed when leaving under . We conclude setting . ∎
Lemma 11.3.
For all exceptional set with respect to , all and all , we have .
Proof.
We fix an exceptional set with respect to , and . For , we write
by the Markov property. But by Lemma 11.2, for all , the law of under is equal to the law of under . Since is exceptional for , we can find properly exceptional for (see Subsection B.1). Hence for all ,
Since has a density by Lemma 11.2, we conclude that and thus that -a.s., we have . All in all, we have proved that , and it suffices to let , since by continuity and since . ∎
Using Lemmas 11.2 and 11.3, it is slightly technical but not difficult to build from and the family a -valued diffusion such that for all , the law of under equals the law of under , for all , setting (so that ), the law of under is the same as that of under and the law of under conditionally on equals the law of under . We have used the notation and .
Remark 11.4.
For all , setting , the law of under is the same as that of under .
Proof.
This follows from Lemma 11.2 when and from the definition of otherwise. ∎
We can finally give the
Proof of Proposition 1.2.
We fix and such that and we prove that defined above is a -process. First, it is clear that is a -process because is a -valued diffusion and since for all , the law of under equals the law of under , with exceptional for . It remains to prove that for all , all and all Lebesgue-null , we have . We set and write, for any ,
Since is -symmetric (because it is a -process), since , where is the semi-group of and since is Lebesgue-null,
Hence there is a Lebesgue-null subset of (depending on ) such that for every . We conclude that
where we finally used Remark 11.4. Since is Lebesgue-null, we deduce from Lemma 11.1 that . Thus , which tends to as because by continuity. ∎
11.2. Final proofs
We fix , such that and a -process , which exists thanks to Subsection 11.1. We recall that was introduced in (2) and define, for all , , as well as the -field
Lemma 11.5.
Fix . If quasi-everywhere, then for all .
Proof.
We fix such that quasi-everywhere. There is an exceptional set such that for all , . We now fix and set . For any ,
By the Markov property and since , we get
But the law of under has a density, so that , whence . Hence and we end with . As usual, we conclude that by letting . ∎
We are now ready to give the
Proof of Proposition 1.3.
Proof of Theorem 1.5.
Let and . Since our -process is a -process, we know from Section 10 that all the conclusions of Theorem 1.5 hold quasi-everywhere. In other words, quasi-everywhere, where is the event on which we have , , there is with cardinal such that there is a -collision in the configuration , etc. We want to prove that for all . By Lemma 11.5, it thus suffices to check that belongs to . But for each , indeed belongs to , because no collision (nor explosion) may happen before getting out of . ∎
We end this section with the following remark (that we will not use anywhere).
Remark 11.6.
Fix and such that . Consider a process and define . For all , there is some -Brownian motion (of dimension ) under such that for all , all ,
(3) |
Proof.
It of course suffices to prove the result during , where . For any and for a given Brownian motion, the solutions to (3) and (1) classically coincide while they remain , because their drift coefficients coincide and are smooth inside . Hence, recalling the notation of Subsection 11.1, it suffices to prove that the semi-groups and of the Markov processes and killed when getting out of coincide for all . By Lemma 11.2, there is an exceptional set such that for all . We next fix . For any , using that has a density and that is Lebesgue-null, we easily deduce that . It is then not difficult, using that is Feller, to let and conclude that indeed, . ∎
Appendix A A few elementary computations
We recall that for and give the
Proof of Lemma 1.1.
First, (3), which says that if and only if , is clear. We next fix , so that and . By concavity of , it only remains to check that (i) , (ii) , and (iii) . We introduce and observe that and that .
For (i), we write since .
For (ii), we have and we need . Writing with an integer and , we need that , and this holds true because and .
For (iii), we write . ∎
We next study the reference measure of the Keller-Segel particle system.
Proposition A.1.
Let and be such that . Recall that and the definition (4) of . (i) The measure is Radon on . (ii) If , then is not Radon on .
Proof.
(i) To show that is radon on , we have to check that for all , which we now fix, there is an open set such that and . We choose , where the balls are subsets of and where
We consider the partition of such that for all in , for all and all , and . Since , it holds that . By definition of and , we see that for all , for all in , for all , all ,
This implies that for some finite constant depending on , for all ,
Recall now that and that we want to show that . Since for all and all , since , and by a translation argument, we are reduced to show that for any , (when , one could study only )
We fix and show that . Since for all , we have , where
But for all ,
by the inequality of arithmetic and geometric means. Thus by symmetry,
Consequently,
Since , we have , so that , whence .
(ii) We next assume that . To prove that is not radon on , we show that for the compact subset
of . All the balls in the previous formula are balls of . For , it holds that are far from each other and far from , which explains that is indeed compact in . There is a positive constant such that for all ,
whence, the value of being allowed to vary,
We now observe that
and that for , we have for all , from which
As a conclusion,
where we finally used the change of variables and . This last integral diverges, because , recall that by definition of . ∎
We need a similar result on the sphere defined in Section 2, where and were also introduced. First, we show an explicit link between and defined in (4) and (1), that we use several times.
Lemma A.2.
We fix , and set . For all Borel ,
Proof.
Since and since is translation invariant,
We next note that is the (true) unit sphere of the -dimensional Euclidean space and proceed to the substitution :
We finally substitute and obtain
But by (4) and , whence
Since finally , the conclusion follows. ∎
We can now study the measure on .
Proposition A.3.
Let and such that . Recall that . (i) The measure is Radon on . (ii) If , then .
Proof.
We start with (i). For , we introduce
Since is compact in , with here the unit ball of , we know from Proposition A.1-(i) that . Now by Lemma A.2,
But for ,
if and only if and . |
Indeed, for all and because and . Thus
All this implies that for all , for almost all , . Since is monotone, we conclude that for all . Since finally and since is compact in for each , we conclude as desired that is Radon on .
We next prove (ii). It holds that , because for , we have . Hence if , then , whence and thus is Radon on by point (i). Since finally is compact, we conclude that . ∎
Appendix B Markov processes and Dirichlet spaces
In a first subsection, we recall some classical definitions and results about Hunt processes, diffusions and Dirichlet spaces found in Fukushima-Oshima-Takeda [11]. In a second subsection, we mention a few results about martingales, times-changes, concatenation, killing and Girsanov transformation of Hunt processes found in [11] and elsewhere.
B.1. Main definitions and properties
Let be a locally compact separable metrizable space endowed with a Radon measure such that Supp . We set , where is a cemetery point. See [11, Section A2] for the definition of a Hunt process : it is a strong Markov process in its canonical filtration, for all , is an absorbing state, i.e. for all under , and a few more technical properties are satisfied. The life-time of is defined by .
Let us denote by its transition kernel. Our Hunt process is said to be -symmetric if for all measurable and all , see [11, page 30]. The Dirichlet space of our Hunt process on is then defined, see [11, page 23], by
The generator of is defined as follows:
and for , we denote by this limit. By [11, Pages 20-21], it holds that
(1) |
and in such a case .
The one-point compactification of is endowed with the topology consisting of all the open sets of and of all the sets of the form with compact in , see page [11, page 69]. Observe that for a -valued sequence , we have if and only if
either , for all large enough, and in the usual sense; or and for all compact subset of , there is such that for all , .
We say that our Hunt process is continuous if is continuous from into , where is endowed with the one-point compactification topology. A continuous Hunt process is called a diffusion.
A Dirichlet space on is said to be regular if it has a core, see [11, page 6], i.e. a subset which is dense in for the norm and dense in for the uniform norm.
Observe two regular Dirichlet spaces and such that for all in a common core are necessarily equal, i.e. and . This follows from the fact that by definition, see [11, page 5], a Dirichlet space is closed.
We say that a Borel set of is -invariant if for all , all we have -a.e, see [11, page 53]. According to [11, page 55], we say that is irreducible if for all -invariant set , we have either or .
We say that is recurrent if for all nonnegative , for -a.e. , we have , see [11, page 55].
We finally say that is transient if for all nonnegative , for -a.e. , we have , with the convention that , see [11, page 55].
By [11, Lemma 1.6.4 page 55], if is irreducible, then it is either recurrent or transient.
A Borel set is properly exceptional if and for all , see [11, page 153]. A property is said to hold true quasi-everywhere if it holds true outside a properly exceptional set.
Remark B.1.
Two Hunt processes with the same Dirichlet space share the same quasi-everywhere notion, up to the restriction that the capacity of every compact set is finite, which is always the case in the present work.
Proof.
We fix a Hunt process and explain why its quasi-everywhere notion depends only on its Dirichlet space. A set is exceptional, see [11, page 152], if there exists a Borel set such that and for -a.e. . A properly exceptional set is clearly exceptional and [11, Theorem 4.1.1 page 155] tells us that any exceptional set is included in a properly exceptional set. Thus, a property is true quasi-everywhere if and only if it holds true outside an exceptional set. Next, [11, Theorem 4.2.1-(ii) page 161] tells us that a set is exceptional if and only if its capacity is , where the capacity of is entirely defined from the Dirichlet space. And for [11, Theorem 4.2.1-(ii) page 161] to apply, one needs that the capacity of all compact sets is finite. ∎
B.2. Toolbox
We start with martingales.
Lemma B.2.
Let be a locally compact separable metrizable space endowed with a Radon measure such that Supp , and a -symmetric -valued diffusion with regular Dirichlet space on and generator . Assume that belongs to and that both and are bounded. Define
with the convention that . Quasi-everywhere, is a -martingale in the canonical filtration of .
This can be found in [11, page 332]. There the assumption on is that there is bounded and measurable such that , i.e. , which simply means that is bounded. Also, the conclusion is that is a MAF, which indeed implies that is a martingale, see [11, page 243].
Next, we deal with time-changes.
Lemma B.3.
Let be a -manifold, a Radon measure on such that , and a -symmetric -valued diffusion with regular Dirichlet space on with core . We also fix continuous and take the convention that . We consider the time-change and its generalized inverse . We introduce . Then is a -symmetric -valued diffusion with regular Dirichlet space on with core , i.e. is the closure of with respect to the norm .
Remark B.4.
If we apply the preceding result to the simple case where is an open subset of and where for all , then when is seen as the Dirichlet form of a -symmetric process, it may be better understood as .
This lemma is nothing but a particular case of [11, Theorem 6.2.1 page 316], see also the few pages before. We only have to check that the Revuz measure in our case is , i.e., see [11, (5.1.13) page 229], that for all bounded nonnegative measurable functions on , for all ,
where is the semi-group of . The left hand side equals , so that the claim is obvious since is -symmetric.
The following concatenation result can be found in Li-Ying [17, Proposition 3.2].
Lemma B.5.
Let be two -manifolds, be some Radon measures on and such that and . Let be a -symmetric -valued diffusion with regular Dirichlet space on with core . Consider , a -symmetric -valued diffusion with regular Dirichlet space on with core . Introduce the measure on . We take the convention that for all , all . Moreover, we set and we define if and . The process
is a -valued -symmetric diffusion , with regular Dirichlet space on with core and, for ,
Observe that may be strictly smaller than due to the identification of all the cemetery points. Also, it actually holds true that on so that the choice is arbitrary but legitimate.
The following killing result is a summary, adapted to our context, of Theorems 4.4.2 page 173 and 4.4.3-(i) page 174 in [11, Section 4.4].
Lemma B.6.
Let be a -manifold, let be a Radon measure on such that , and let be a -symmetric -valued diffusion with regular Dirichlet space on with core . Let be an open subset of and consider , with the convention that . Then, setting
is a -symmetric -valued diffusion with regular Dirichlet space on with core and for ,
Note that since is an open subset of the manifold and since the Hunt process is continuous, the regularity condition (4.4.6) of [11, Theorem 4.4.2 page 173] is obviously satisfied.
We finally give an adaptation of the Girsanov theorem in the context of Dirichlet spaces, which is a particular case of Chen-Zhang [5, Theorem 3.4].
Lemma B.7.
Let be an open subset of , with , be a Radon measure on such that and be a -symmetric -valued diffusion with regular Dirichlet space on with core such that for all ,
Let stand for its generator. Let be bounded, such that for , we have with is bounded. Set
with the conventions that and . Assume that is continuous on . Then quasi-everywhere, is a bounded -martingale under , where we have set , and there exists a probability measure on , such that for all , on .
Moreover is a -symmetric -valued diffusion with regular Dirichlet space on such that for all ,
Actually, they speak of right processes in [5], but this is not an issue since we only consider continuous Hunt processes. Also, they assume that is bounded from above and from below by some deterministic constants, on each compact time interval, but this is obvious under our assumptions on and . Finally, their expression of is different, see [5, pages 485-486]: first, they define as the martingale part of . By Lemma B.2 (applied to ), we see that
Then they put and define as
But by Itô’s formula, , whence , so that as desired.
References
- [1] G. Arumugam, J. Tyagi, Keller-Segel Chemotaxis Models: A Review, Acta Appl. Math. 171 (2021), 6.
- [2] A. Blanchet, J. Dolbeault, B. Perthame, Two-dimensional Keller-Segel model: optimal critical mass and qualitative properties of the solutions, Electron. J. Differential Equations 44 (2006), 32 pp.
- [3] D. Bresch, P.E. Jabin, Z. Wang On mean-field limits and quantitative estimates with a large class of singular kernels: application to the Patlak-Keller-Segel model, C. R. Math. Acad. Sci. Paris 357 (2019), 708–720.
- [4] P. Cattiaux, L. Pédèches, The 2-D stochastic Keller-Segel particle model: existence and uniqueness, ALEA, Lat. Am. J. Probab. Math. Stat. 13 (2016), 447–463.
- [5] Z. Chen, T.S. Zhang, Girsanov and Feynman-Kac type transformations for symmetric Markov processes. Ann. Inst. H. Poincaré Probab. Statist. 38 (2002), 475–505.
- [6] J. Dolbeault, C. Schmeiser, The two-dimensional Keller-Segel model after blow-up, Discrete Contin. Dyn. Syst. 25 (2009), 109–121.
- [7] I. Fatkullin, A study of blow-ups in the Keller-Segel model of chemotaxis, Nonlinearity 26 (2013), 81–94.
- [8] N. Fournier, B. Jourdain, Stochastic particle approximation of the Keller-Segel equation and two-dimensional generalization of Bessel processes, Ann. Appl. Probab. 27 (2017), 2807–2861.
- [9] N. Fournier, M. Hauray, S. Mischler, Propagation of chaos for the 2D viscous vortex model, J. Eur. Math. Soc. 16 (2014), 1423–1466.
- [10] M. Fukushima, From one dimensional diffusions to symmetric Markov processes, Stochastic Process. Appl. 120 (2010), 590–604.
- [11] M. Fukushima, Y. Oshima, M. Takeda, Dirichlet forms and symmetric Markov processes. Second revised and extended edition. Walter de Gruyter, 2011.
- [12] J. Haškovec, C. Schmeiser, Stochastic particle approximation for measure valued solutions of the 2D Keller-Segel system, J. Stat. Phys. 135, 133–151.
- [13] J. Haškovec, C. Schmeiser, Convergence of a stochastic particle approximation for measure solutions of the 2D Keller-Segel system, Comm. Partial Differential Equations 36(6) (2011), 940–960.
- [14] J.F. Jabir, D. Talay, M. Tomasevic, Mean-field limit of a particle approximation of the one-dimensional parabolic-parabolic Keller-Segel model without smoothing, Electron. Commun. Probab. 23 (2018), Paper No. 84.
- [15] I. Karatzas, S.E. Shreve, Brownian motion and stochastic calculus. Second edition. Graduate Texts in Mathematics, 113. Springer-Verlag, New York, 1991.
- [16] E.F. Keller, L.A. Segel, Initiation of slime mold aggregation viewed as an instability, J. Theor. Biol. 26 (1970), 399–415.
- [17] L. Li,, J. Ying, Regular subspaces of Dirichlet forms. Festschrift Masatoshi Fukushima, 397–420, Interdiscip. Math. Sci., 17, World Sci. Publ., Hackensack, NJ, 2015.
- [18] C. Olivera, A. Richard, M. Tomasevic, Particle approximation of the 2-d parabolic-elliptic Keller-Segel system in the subcritical regime, arXiv:2011.00537.
- [19] H. Osada, Propagation of chaos for the two-dimensional Navier-Stokes equation, Proc. Japan Acad. Ser. A Math. Sci. 62 (1986), 8–11.
- [20] C.S. Patlak, Random walk with persistence and external bias, Bull. Math. Biophys. 15 (1953), 311–338.
- [21] D. Revuz, M. Yor, Continuous martingales and Brownian motion. Third Edition. Springer, 2005.
- [22] A. Stevens, The derivation of chemotaxis equations as limit dynamics of moderately interacting stochastic many-particle systems, SIAM J. Appl. Math. 61 (2000), 183–212.
- [23] T. Suzuki Exclusion of boundary blowup for 2D chemotaxis system provided with Dirichlet boundary condition for the Poisson part, J. Math. Pures Appl. 100 (2013), 347–367.
- [24] J.J.L. Velazquez, Point dynamics in a singular limit of the Keller-Segel model. I. Motion of the concentration regions, SIAM J. Appl. Math., 64 (2004), 1198–1223.
- [25] J.J.L. Velazquez, Point dynamics in a singular limit of the Keller-Segel model. II. Formation of the concentration regions, SIAM J. Appl. Math., 64 (2004), 1224–1248.